Você está na página 1de 7

Journal of Chromatography B, 878 (2010) 1257–1263

Contents lists available at ScienceDirect

Journal of Chromatography B
journal homepage: www.elsevier.com/locate/chromb

Short communication

Revised method for routine determination of urinary dialkyl phosphates using


gas chromatography–mass spectrometry夽
Jun Ueyama a,∗ , Michihiro Kamijima b , Takaaki Kondo a , Kenji Takagi a , Eiji Shibata c ,
Takaaki Hasegawa d , Shinya Wakusawa a , Tomoko Taki e , Masahiro Gotoh f , Isao Saito g
a
Program in Radiological and Medical Laboratory Sciences, Nagoya University Graduate School of Medicine, Nagoya, Japan
b
Department of Occupational and Environmental Health, Nagoya City University Graduate School of Medical Sciences, Nagoya, Japan
c
Department of Health and Psychosocial Medicine, Aichi Medical University School of Medicine, Aichi, Japan
d
Department of Hospital Pharmacy and Pharmacokinetics, Aichi Medical University School of Medicine, Aichi, Japan
e
COOP Aichi, Nagoya, Japan
f
Department of Occupational and Environmental Health, Nagoya University Graduate School of Medicine, Nagoya, Japan
g
Food Safety and Quality Research Center, Tokai COOP Federation, Aichi, Japan

a r t i c l e i n f o a b s t r a c t

Article history: Among urinary organophosphorus pesticide (OP) metabolites, dialkyl phosphates (DAPs) have been
Received 4 September 2009 most often measured as a sensitive biomarker in non-occupational and occupational OP exposure risk
Accepted 8 February 2010 assessment. In our conventional method, we have employed a procedure including simple liquid–liquid
Available online 12 February 2010
extraction (diethyl ether/acetonitrile), derivatization (pentafluorobenzylbromide, PFBBr) and clean-up
(multi-layer column) for gas chromatography–mass spectrometry (GC–MS) analysis starting from 5-mL
Keywords:
urine samples. In this study, we introduce a revised analytical method for urinary DAPs; its main mod-
Organophosphorus insecticide
ification was aimed at improving the pre-derivatization dehydration procedure. The limits of detection
Urinary metabolite
Dialkylphosphate
were approximately 0.15 ␮g/L for dimethylphosphate (DMP), 0.07 ␮g/L for diethylphosphate (DEP), and
GC–MS 0.05 ␮g/L for both dimethylthiophosphate (DMTP) and diethylthiophosphate (DETP) in 2.5-mL human
Biological monitoring urine samples. Within-run precision (percent of relative standard deviation, %RSD) at the DAP levels
varying in the range of 0.5–50 ␮g/L was 6.0–19.1% for DMP, 3.6–18.3% for DEP, 8.0–25.6% for DMTP
and 9.6–27.8% for DETP. Between-run precision at 5 ␮g/L was below 15.7% for all DAPs. The revised
method proved to be feasible to routine biological monitoring not only for occupational OP exposure but
also for environmental background levels in the general population. Compared to our previous method,
the revised method underscores the importance of adding pre-derivatization anhydration for higher
sensitivity and precision.
© 2010 Elsevier B.V. All rights reserved.

1. Introduction Environmental Protection Agency (EPA)-registered OPs are metab-


olized to dialkyl phosphates (DAPs), including dimethylphosphate
Organophosphorus compounds (OPs) have been widely and (DMP), diethylphosphate (DEP), dimethylthiophosphate (DMTP),
effectively used as insecticides with applications in agricultural diethylthiophosphate (DETP), dimethyldithiophosphate (DMDTP),
settings, public health, commerce, and individual households and diethyldithiophosphate (DEDTP) [2,3], these DAPs in urine have
throughout the world [1]. Growing concern over the long- been measured as biomarkers of OP exposure [4,5]. Development
term effects of low-level exposure to OPs on human health of analytical equipment and protocols of sample preparation has
has encouraged more detailed research both in experimental made it possible to detect low-level DAPs from various general
and in epidemiological settings. Since about 75% of the US populations, revealing that OP exposure has commonly occurred
even in ordinary daily life. However, the main pathway of human
exposure (e.g. ingestion, inhalation or dermal absorption) as well as
toxicity due to long-term exposure to low-level doses still remains
夽 This paper is part of the special issue ‘Bioanalysis of Organophosphorus Toxicants to be explored. This is partly because of the difficulty in deter-
and Corresponding Antidotes’, Harald John and Horst Thiermann (Guest Editors). mining low-level DAPs among biological samples obtained from
∗ Corresponding author at: Department of Medical Technology, School of Health
the general population, due to the complicated and costly sam-
Sciences, Nagoya University, 1-1-20 Daikominami, Higashi-ku, Nagoya, Aichi 461-
8673, Japan. Tel.: +81 52 719 1341; fax: +81 52 719 1341. ple preparations, overall low-throughput yield, and requirement of
E-mail address: ueyama@met.nagoya-u.ac.jp (J. Ueyama). expensive equipment. Urine sample has been frequently used for

1570-0232/$ – see front matter © 2010 Elsevier B.V. All rights reserved.
doi:10.1016/j.jchromb.2010.02.005
1258 J. Ueyama et al. / J. Chromatogr. B 878 (2010) 1257–1263

the determination of DAP because they are collected non-invasively before collection, were used for the basic methodological exami-
and contain DAPs in higher concentrations than other biological nation in this study.
samples [6,7]. A flow chart of the urinary DAPs determination procedure is
We previously reported a method for DAP analysis [8], which shown in Fig. 1. Urine sample (2.5 mL) was pipetted into a 10-
is relatively simple and sensitive enough to be adopted in routine mL screw-top glass test tube, and 20 ␮L of I.S. solution (100 mg/L
biological monitoring of non-occupational as well as occupational DBP), 2.5 g of NaCl, 1 mL of 6 M HCl, 50 mg of Na2 S2 O5 and 2.5 mL
exposure to OPs. However, there were still some shortcomings to of diethylether–acetonitrile (1:1, v/v) were added. After vigor-
be resolved such as low precision and the need for a relatively ous mechanical shaking for 5 min, the test tube was centrifuged
large volume of urine (>5 mL). The aim of the present study was to (1500 × g for 5 min at room temperature). The organic phase
improve our previous methods [8] for urinary DAP measurement (upper layer) containing DAPs was passed through 1 g of anhy-
using gas chromatography–mass spectrometry (GC–MS) equipped drous sodium sulfate column and collected into a new screw-top
with an electron ionization system and achieve higher sensitivity glass test tube containing 15 mg of K2 CO3 . The aqueous phase
and precision with high throughput and lower cost. was re-extracted with 2.5 mL of diethylether–acetonitrile (1:1, v/v)
and then centrifuged. The supernatant obtained from the second
2. Experimental extraction was passed through the anhydrous sodium sulfate col-
umn, and combined with the first extract. The resulting extract was
2.1. Reagents evaporated at 40 ◦ C (heat block) to dryness with a gentle nitrogen
stream for about 1 h. To the dried extracts, 15 mg of K2 CO3 , 1 mL
DMP tetramethylammonium salt (99.9% purity), DMTP ammo- of acetonitrile and 10 ␮L of PFBBr were added and incubated in a
nium salt (98.9%), DEP (98.2%) and DETP ammonium salt (95.2%) heat block at 80 ◦ C for 30 min with occasional swirling. Afterwards,
were obtained from Hayashi Pure Chemical Ind. (Osaka, Japan), and 3 mL of water and 3 mL of n-hexane were added, and the mixture
dibutylphosphate (DBP), used for an internal standard (I.S.), was was shaken vigorously for 5 min and centrifuged for 5 min (1500 × g
from Tokyo Kasei Kogyo (Tokyo, Japan). Diethyl ether, acetonitrile, for 5 min). The upper layer containing PFB-DAPs was transferred to
n-hexane, acetone and toluene, which are pesticide residue grade, new test-glass tubes. The extraction was then repeated with 3 mL
and sodium sulfate anhydrous, sodium chloride (NaCl), sodium of n-hexane, and the supernatant obtained from the second extrac-
disulfite (Na2 S2 O5 ) and 6 M hydrochloric acid, were purchased tion was combined with the first extract. The combined extract was
from Kanto Chemicals (Tokyo, Japan). Pentafluorobenzylbromide loaded into a clean-up column, followed by washing with 5 mL of
(PFBBr) was purchased from Sigma–Aldrich (St. Louis, MO, USA), acetone-n-hexane (2:98, v/v) for removing unreacted PFBBr. PFB-
Florisil (60–100 mesh) from Wako Pure Chemicals (Osaka, Japan), DAPs were then eluted with 5 mL of acetone-n-hexane (15:85, v/v),
and Primary/Secondary Amine (PSA) 40 ␮m from Varian, Inc. (Palo and the eluate was evaporated at 45 ◦ C to dryness with a gentle
Alto, CA, USA). Water used throughout the experiments was dis- nitrogen stream for about 15 min. The residue was dissolved in
tilled and deionized to 18 M with a Millipore Milli-Q System 200 ␮L of toluene and injected into GC–MS.
(Millipore Co., Bedford, MA, USA). All other reagents were of
analytical grade purity. Muromac Mini Column (medium size, 2.4. Assay validation
110 mm × 8.0 mm i.d.) (Muromachi Chemical Inc., Fukuoka, Japan)
was used for dehydration and clean-up process. The clean-up col- Using the proposed method, two calibration curves were sepa-
umn could be readily prepared without any special technique and rately prepared using pooled urine. The first curve corresponded to
was composed of 0.3 g of Florisil (lower), 0.1 g of PSA (middle) and the concentrations of urinary DAPs ranging from 0.5 to 50 ␮g/L (four
0.5 g of sodium sulfate anhydrous (upper). points), and the second curve to the range from 50 to 1000 ␮g/L
(five points); the latter curve was used along with the former one
to determine the urinary DEP, DMTP and DETP in occupational OP
2.2. Apparatus and GC–MS conditions exposures.
To determine and calculate absolute recoveries, we spiked DAPs
Analyses of DAPs derivatized with pentafluorobenzylbromide at two different stages in the procedure; i.e. in the beginning of the
were performed using an Agilent 5975 inert MSD system. The GC extraction procedure (urine sample) and prior to the derivatization
operating conditions were as follows: GC column, Rtx-65 (Restek, procedure. We compared the I.S. ratios obtained at these two stages.
USA), 30 m × 0.25 mm i.d., 0.25-␮m film thickness; column tem- Calibration curves were represented by the analyte/I.S. peaks
peratures, 70 ◦ C (1 min)–15 ◦ C/min–300 ◦ C (6 min); injection port area ratio versus the concentrations of the calibration samples.
temperature, 250 ◦ C; carrier gas, helium (99.999% purity); flow rate, The within-run precision for our revised method was examined
1 mL/min. The injection volume was 1 ␮L. Splitless was changed through the assay of the pooled urine spiked with DAP concentra-
to split 15:1 at 2 min after sample injection. The MS operating tions of 0.5, 2.5, 5 and 50 ␮g/L (n = 4–5). Moreover, the between-run
conditions were as follows: ionization source temperature, 230 ◦ C; precision was examined through the duplicate assay of the pooled
electron ionization, 70 eV; interface temperature, 300 ◦ C; injection urine spiked with DAPs at a concentration of 1, 5 and 500 ␮g/L for 4
pressure, 88 psi. Chromatogram peak was identified by target and consecutive days (n = 3–4). The limits of detection (LOD) and limit
quantifier ions for each pentafluorobenzyl (PFB)-DAP as shown in of quantitation (LOQ) were calculated based on the signal-to-noise
Table 1. Use of C- and Q-ion presented in this table is appropriate ratio of 3 and 10, respectively.
for selectivity and sensitivity under new analytical conditions.
2.5. Application of methods to field study samples
2.3. Standard preparation and analytical procedure
DAPs were measured in morning urine samples collected
Each standard (DMP, DEP, DMTP and DETP) was prepared from 25 healthy OP non-exposed volunteers aged 40 ± 9 years
at a concentration of 1000 mg/L in methanol, and diluted with (mean ± S.D.) in February 2009 (low-exposure urine). The meth-
methanol to each working standard solution. The standard solu- ods were also applied to morning urine samples collected from 25
tions were stored at 4 ◦ C in the dark, and were used within 1 month OP exposed persons aged 37 ± 9 years (mean ± S.D.) in August 2009
of their preparation. Urine samples from healthy volunteers, who (high-exposure urine); they were all workers engaged in pest con-
were neither treated with any drugs nor exposed to chemicals trol occupation (PCO) located in the Chubu area (central Japan) and
J. Ueyama et al. / J. Chromatogr. B 878 (2010) 1257–1263 1259

Table 1
Chemical structures, fragment ions and retention time of dialkyl phosphates and dibutyl phosphate.

Compound Structure m/za Retention time (min)


b c
C-ion Q-ion

DMP 110 306 8.86

194
DEP 258 334 9.44

197
DMTP 322 322 10.28

211
DETP 213 350 10.73

274

DBP (internal standard) 335 335 11.52

a
m/z: mass/charge ratio.
b
C-ion: selected ions for confirmation.
c
Q-ion: selected ions for quantification.

asked to provide urine samples collected on the day after insecti- DETP concentrations in the general population. Absolute recov-
cide spraying. The reason for the difference in the sample collection ery data are approximately similar to those of our previous report
season was to verify whether the present method is feasible for a [8], and superior to the solid-phase extraction recovery except for
wide range of DAP concentrations; pesticide exposure is lowest in DEP [14]. The mean absolute recoveries for DMP were lower than
the general population in winter [9] and highest in PCOs in sum- those for other DAPs due to its high polarity. This lower recovery
mer [10]. Collected urine samples were immediately transferred might increase the LOD value for DMP. Because absolute recover-
into 10 mL polyethylene tubes and stored at −80 ◦ C until DAP assay. ies of DAPs in water differ from those in urine matrix (data not
The Ethics Committee of the Nagoya University Graduate School of shown), pooled urine should be used for a matrix matched cali-
Medicine and Nagoya City University Graduate School of Medical bration curve. LOD and LOQ values are below or similar to those
Sciences, Nagoya, Japan approved the study protocol. When uri- reported by De Alwis et al. [14,15], who used solid-phase extrac-
nary DAP concentrations were less than LOD, they were estimated tion and GC–MS/MS, Dulaurent et al. [16], who used liquid–liquid
as half the LOD value for statistical analyses [11]. extraction and LC–MS/MS, or by Ueyama et al. [8]. Concentrations
of LOQ are lower than the geometric mean of urinary DAPs in
3. Results and discussion the general populations including children and pregnant women
[12,17–19], suggesting that our present method is sensitive enough
3.1. Assay validation for monitoring urinary DAPs in most general populations. The ratios
of C- and Q-ion abundances for each DAP were almost invariable at
The validation parameters are summarized in Table 2. For concentrations both above and below LOD. It is unlikely that some
the within-run precision, percent of relative standard deviation interference substances deteriorate selectivity for urinary DAP con-
(%RSD) ranged from 5.6 to 27.8% for all DAPs. For the between-run centrations below LOD.
precision, the %RSD was between 7.0 and 51.3%. Reproducibility
deteriorated when the analyte/I.S. peak area ratio was not used 3.2. Advantages of this method
(data not shown). Relatively high %RSD values were shown for
the within-run precision at 0.5 ␮g/L (DMTP 25.6% and DETP 27.8%) Advantages of the present method over the previous one were
and between-run precision at 1 ␮g/L (DMP 36.7%, DMTP 51.3% and smaller sample volume required without compromising the high
DETP 23.0%). The urinary DETP concentration of 0.5 ␮g/L is near sensitivity and precision, cost-effectiveness and smaller volume of
the geometric mean in the general population [5,12]. On the other such reagents as diethylether/acetonitrile mixture, NaCl and irri-
hand, geometric means of urinary DMP and DMTP level in the tating nature PFBBr.
general population are around 10 ␮g/L or more [12,13]. Therefore, Previously, urinary DAPs have been extracted by various
biomonitoring of DETP at low concentrations is less precise in the methods: liquid–liquid extraction [8,20,21], solid-phase extrac-
assessment of the OP exposure level for the general population. tion [14,15,22,23], and lyophilization [24]. Weerasekera et al. [3]
Further studies are needed to accurately determine the urinary suggested that solid-phase extraction (ChemElute cartridge) and
1260 J. Ueyama et al. / J. Chromatogr. B 878 (2010) 1257–1263

Fig. 1. Analytical procedure for urinary DAPs.

analysis using the GC–MS is the most cost-effective and rapid extraction as a more cost-effective and highly sensitive DAP extrac-
method. There were some drawbacks to other extraction meth- tion procedure. We hypothesized that the derivatization procedure
ods; lyophilization is time-consuming and liquid–liquid extraction using PFBBr would be adversely affected by any interference sub-
is less accurate and precise [3]. If the accuracy and precision are stance in extracts after the liquid–liquid extraction procedure.
remediated, it should be possible to develop the liquid–liquid It is well known that the derivatization reaction of PFBBr is
J. Ueyama et al. / J. Chromatogr. B 878 (2010) 1257–1263 1261

Table 2
Accuracy, precision, LOD and LOQ data of analytical procedure.

Pooled urine spiked concentration (␮g/L of urine) na DMP DEP DMTP DETP

Within-run
Precision (%RSDb ) 0.5 5 19.1 18.3 25.6 27.8
2.5 4 7.6 6.3 8.0 12.3
5 4 5.6 4.1 12.5 15.1
50 4 6.0 3.6 15.6 9.6
Mean recoverye (%) 1 4 62.6 87.9 84.0 102.0
50 4 68.6 85.6 88.4 97.7
500 4 71.3 90.5 83.5 91.6

Between-run
Precision (%RSD) 1 3 36.7 16.4 51.3 23.0
5 4 15.7 10.3 9.3 7.0
500 3 10.2 15.1 9.9 11.3

R2 of calibration line
0.5–50 ␮g/L of urine 0.995 0.998 0.984 0.974
50–1000 ␮g/L of urine 0.973 0.980 0.991 0.989

LODc (␮g/L) (signal-to-noise ratio = 3) 0.15 0.07 0.05 0.05


LOQd (␮g/L) (signal-to-noise ratio = 10) 0.5 0.3 0.2 0.2
a
n: number of observations.
b
RSD: relative standard deviation.
c
LOD: limit of detection.
d
LOQ: limit of quantitation.
e
Recovery given by adding the standards on derivatization step.

inhibited in the presence of water. In the extraction procedure, tions at 5 ␮g/L in the present method (5.6 for DMP and 4.1 for DEP)
the organic phase obtained from liquid–liquid extraction using were lower than those in our previous method (16.1 for DMP and
diethylether–acetonitrile (1:1, v/v) may contain much water. The 20.6 for DEP) [8]. The LOD values of DAPs reported in our pre-
water retained in high hydroscopic solid K2 CO3 could have affected vious method [8] were decreased by half in the present method.
the efficiency of derivatizing DAPs with PFBBr, thus resulting in The enhanced sensitivity is likely due to improved derivatization
the deteriorated validation data even if the post-derivatization efficiency. While Timchalk et al. [25] adopted dehydration proce-
dry up procedure was conducted. Therefore, we examined how dure for DAP measurement in rat biological sample, the present
the dehydration procedure prior to PFBBr derivatization affects study is, to our knowledge, the first to indicate that the dehydration
the accuracy, precision and sensitivity. The improvement found in procedure prior to derivatization can improve sensitivity, accuracy
our validation data is mainly attributable to the pre-derivatization and precision for the measurement of human urinary DAPs using
dehydration. For example, within-run precision (RSD%) concentra- liquid–liquid extraction and PFBBr derivatization. Moreover, the

Fig. 2. Typical SIM GC–MS chromatograms in human urine samples under the LOD and near the LOQ level (A, DMP 1.0 ␮g/L; B, DEP 0.4 ␮g/L; C, DMTP 1.0 ␮g/L; D, DETP
0.2 ␮g/L), and DAP-spiked urine (E, DMP 47 ␮g/L; F, DEP 35 ␮g/L; G, DMTP 78 ␮g/L; H, DETP 22 ␮g/L). Detected masses for quantification were m/z 306 for DMP, 334 for DEP,
322 for DMTP, and 350 for DETP.
1262 J. Ueyama et al. / J. Chromatogr. B 878 (2010) 1257–1263

Table 3
Level of dialkyl phosphates in morning urine from healthy volunteers (control, n = 25) and persons engaged in pest control occupation (PCO, n = 25).

Dialkyl phosphates Detected (%) Geometric mean (␮g/L of urine) Median (␮g/L of urine) 95th percentile (␮g/L of urine) Maximum (␮g/L of urine)

DMP
Controls 88 4.2 8.0 41.3 43.7
PCOs 100 70.0 39.1 765.9 862.0

DEP
Controls 100 1.3 1.1 15.8 20.1
PCOs 100 7.1 5.8 92.3 106.8

DMTP
Controls 100 4.2 3.9 40.7 45.9
PCOs 100 36.0 38.5 914.9 1044.2

DETP
Controls 88 0.2 0.2 5.3 6.5
PCOs 100 0.9 0.7 26.1 28.0

clean-up procedure after derivatization using a multi-layer column are needed to establish a feasible and scientifically acceptable data
effectively removed the highly irritating PFBBr and other interfer- collection method to reflect the 24-h total excretion level.
ing substances, resulting in a drastic reduction of unidentifiable Our method can be applied for DAP measurement in other bio-
peaks in both the total ion chromatogram (TIC) and the selected ion logical samples such as blood, hair and amniotic fluid with slight
monitoring (SIM) chromatograms, and stabilization of the baseline. modifications. Previously, Margariti et al. [28] determined hair
Both GC–MS and LC–MS/MS methods have been developed to DAPs using the same technique as ours with slight modifications.
detect DAPs in urine [11,15,16], and these sophisticated analyti-
cal tools have inherent advantages in terms of high selectivity and 4. Conclusion
sensitivity. The major advantage of the LC–MS/MS-based method
is its ability to analyze various metabolites simultaneously with Pre-derivatization process for effective anhydration, combined
substantially simple sample pretreatment. However, these advan- with a present GC–MS system, clearly improved sensitivity, pre-
tages are partially counterbalanced by the higher instrument cost cision, stability, and throughput without additional high-cost
in comparison with GC–MS systems. Moreover, sensitivity in DAP requirements. The present method would allow many laborato-
measurement using LC–MS/MS [15,16] at present tends to be ries to conduct the routine biological monitoring of urinary DAPs
lower than that using GC–MS or GC–MS/MS method [8,14,15]. In in the general population, and will be helpful in epidemiological
this study we combined the liquid–liquid extraction and GC–MS study dealing with possible toxicity from low-level, long-term OP
determination to achieve a lower-cost, faster and higher sample exposure.
processing capacity.
Acknowledgement

3.3. Application of methods to field study samples This work was supported in part by a Grant-in-Aid for Scientific
Research and a Grant-in-Aid for Young Scientists from the Japan
Fig. 2 shows the typical chromatogram in the SIM mode of under Society for the Promotion of Science (JSPS).
the LOD, near the LOQ and DAP spiked urine. Urinary concentra-
tion of DAPs in 25 controls and 25 PCOs are summarized in Table 3.
References
DAPs were detected in 88–100% of all human samples. Geomet-
ric mean, median, 95th percentile and maximum levels of each [1] P.G. Bardin, S.F. van Eeden, J.A. Moolman, A.P. Foden, J.R. Joubert, Arch. Intern.
DAP in the exposed group were obviously higher than those in the Med. 154 (1994) 1433.
[2] R. Bravo, W.J. Driskell, R.D. Whitehead Jr., L.L. Needham, D.B. Barr, J. Anal. Tox-
non-exposed one. The urinary DAP concentrations in the exposed
icol. 26 (2002) 245.
workers are higher than those in our previous report [8]. The differ- [3] G. Weerasekera, K.D. Smith, L.L. Needham, D.B. Barr, J. Anal. Toxicol. 32 (2008)
ence in the urine collection method may be the reason for this. Urine 106.
samples were collected on the next day after insecticide spraying in [4] C. Lu, K. Toepel, R. Irish, R.A. Fenske, D.B. Barr, R. Bravo, Environ. Health Perspect.
114 (2006) 260.
this study. But in our previous study, urine samples were collected [5] M. Valcke, O. Samuel, M. Bouchard, P. Dumas, D. Belleville, C. Tremblay, Int.
when the workers underwent the health checkup. The urinary DAP Arch. Occup. Environ. Health 79 (2006) 568.
concentrations in the non-exposed persons were approximately [6] A. Bradman, D.B. Barr, B.G. Claus Henn, T. Drumheller, C. Curry, B. Eskenazi,
Environ. Health Perspect. 111 (2003) 1779.
the same or less than those in previous reports [12,13,17,18]. [7] R.M. Whyatt, D.B. Barr, Environ. Health Perspect. 109 (2001) 417.
One disadvantage in monitoring urinary DAP is the difficulty [8] J. Ueyama, I. Saito, M. Kamijima, T. Nakajima, M. Gotoh, T. Suzuki, E. Shibata,
in estimating the OP exposure level using the measurement of T. Kondo, K. Takagi, K. Miyamoto, J. Takamatsu, T. Hasegawa, J. Chromatogr. B
832 (2006) 58.
DAPs from biological samples. DAP levels determined from bio- [9] G.S. Berkowitz, J. Obel, E. Deych, R. Lapinski, J. Godbold, Z. Liu, P.J. Landrigan,
logical sample might be affected by the intake of environmental M.S. Wolff, Environ. Health Perspect. 111 (2003) 79.
DAP residue, thereby resulting in overestimation of the predicted [10] D. Wang, M. Kamijima, R. Imai, T. Suzuki, Y. Kameda, K. Asai, A. Okamura, H.
Naito, J. Ueyama, I. Saito, T. Nakajima, M. Goto, E. Shibata, T. Kondo, K. Takagi,
dose of OPs. In fact, some researchers have reported the existence
S. Wakusawa, J. Occup. Health 49 (2007) 509.
of DAPs in many foods [26]. Urinary DAP levels are usually reported [11] M.M. Finkelstein, D.K. Verma, AIHAJ 62 (2001) 195.
as volume-weighted concentrations (e.g., ␮g/L) or creatinine- [12] K. Becker, M. Seiwert, J. Angerer, M. Kolossa-Gehring, H.W. Hoppe, M. Ball, C.
Schulz, J. Thumulla, B. Seifert, Int. J. Hyg. Environ. Health 209 (2006) 221.
adjusted concentrations (e.g., ␮g/g creatinine). But the amount of
[13] U. Heudorf, J. Angerer, Environ. Res. 86 (2001) 80.
daily creatinine excretion in urine varies according to age, sex, mus- [14] G.K.H. De Alwis, L.L. Needham, D.B. Barr, J. Anal. Toxicol. 32 (2008)
cle mass and diet. Fortin et al. [27] suggested that the measurement 721.
of insecticide metabolite from spot urine samples may lead to seri- [15] G.K.H. De Alwis, L.L. Needham, D.B. Barr, J. Chromatogr. B 843 (2006)
34.
ous errors in the estimation of the actual daily absorbed doses, [16] S. Dulaurent, F. Saint-Marcoux, P. Marquet, G. Lachatre, J. Chromatogr. B 831
even with adjustment of the creatinine contents. Further studies (2006) 223.
J. Ueyama et al. / J. Chromatogr. B 878 (2010) 1257–1263 1263

[17] X. Ye, F.H. Pierik, R. Hauser, S. Duty, J. Angerer, M.M. Park, A. Burdorf, A. Hof- [22] W.C. Lin, C.H. Kuei, H.C. Wu, C.C. Yang, H.Y. Chang, J. Anal. Toxicol. 26 (2002)
man, V.W. Jaddoe, J.P. Mackenbach, E.A. Steegers, H. Tiemeier, M.P. Longnecker, 176.
Environ. Res. 108 (2008) 260. [23] C. Aprea, G. Sciarra, L. Lunghini, J. Anal. Toxicol. 20 (1996) 559.
[18] C. Saieva, C. Aprea, R. Tumino, G. Masala, S. Salvini, G. Frasca, M.C. Giur- [24] A.N. Oglobline, H. Elimelakh, B. Tattam, R. Geyer, G.E. O’Donnell, G. Holder,
danella, I. Zanna, A. Decarli, G. Sciarra, D. Palli, Sci. Total Environ. 332 (2004) Analyst 126 (2001) 1037.
71. [25] C. Timchalk, A. Busby, J.A. Campbell, L.L. Needham, D.B. Barr, Toxicology 237
[19] C. Aprea, M. Strambi, M.T. Novelli, L. Lunghini, N. Bozzi, Environ. Health Per- (2007) 145.
spect. 108 (2000) 521. [26] X. Zhang, J.H. Driver, Y. Li, J.H. Ross, R.I. Krieger, J. Agric. Food Chem. 56 (2008)
[20] T.F. Moate, C. Lu, R.A. Fenske, R.M. Hahne, D.A. Kalman, J. Anal. Toxicol. 23 (1999) 10638.
230. [27] M.C. Fortin, G. Carrier, M. Bouchard, Environ. Health 7 (2008) 55.
[21] J. Hardt, J. Angerer, J. Anal. Toxicol. 24 (2000) 678. [28] M.G. Margariti, A.M. Tsatsakis, Biomarkers 14 (2009) 137.

Você também pode gostar